Preparation and Analysis of Linguistic Corporaide/papers/Humbook.pdfPreparation and Analysis of...

32
Preparation and Analysis of Linguistic Corpora The corpus is a fundamental tool for any type of research on language. The availability of computers in the 1950’s immediately led to the creation of corpora in electronic form that could be searched automatically for a variety of language features and compute frequency, distributional characteristics, and other descriptive statistics. Corpora of literary works were compiled to enable stylistic analyses and authorship studies, and corpora representing general language use became widely used in the field of lexicography. In this era, the creation of an electronic corpus required entering the material by hand, and the storage capacity and speed of computers available at the time put limits on how much data could realistically be analyzed at any one time. Without the Internet to foster data sharing, corpora were typically created, and processed at a single location. Two notable exceptions are the Brown Corpus of American English (Francis and Kucera, 1967) and the London/Oslo/Bergen (LOB) corpus of British English (Johanssen et al., 1978); both of these corpora, each containing one millions words of data tagged for part of speech, were compiled in the 1960’s using a representative sample of texts produced in the year 1961. For several years, the Brown and LOB were the only widely available computer-readable corpora of general language, and therefore provided the data for numerous language studies. In the 1980’s, the speed and capacity of computers increased dramatically, and, with more and more texts being produced in computerized form, it became possible to create corpora much larger than the Brown and LOB, containing millions of words. The availability of language samples of this magnitude opened up the possibility of gathering

Transcript of Preparation and Analysis of Linguistic Corporaide/papers/Humbook.pdfPreparation and Analysis of...

Preparation and Analysis of Linguistic Corpora

The corpus is a fundamental tool for any type of research on language. The availability of

computers in the 1950’s immediately led to the creation of corpora in electronic form that

could be searched automatically for a variety of language features and compute

frequency, distributional characteristics, and other descriptive statistics. Corpora of

literary works were compiled to enable stylistic analyses and authorship studies, and

corpora representing general language use became widely used in the field of

lexicography. In this era, the creation of an electronic corpus required entering the

material by hand, and the storage capacity and speed of computers available at the time

put limits on how much data could realistically be analyzed at any one time. Without the

Internet to foster data sharing, corpora were typically created, and processed at a single

location. Two notable exceptions are the Brown Corpus of American English (Francis

and Kucera, 1967) and the London/Oslo/Bergen (LOB) corpus of British English

(Johanssen et al., 1978); both of these corpora, each containing one millions words of

data tagged for part of speech, were compiled in the 1960’s using a representative sample

of texts produced in the year 1961. For several years, the Brown and LOB were the only

widely available computer-readable corpora of general language, and therefore provided

the data for numerous language studies.

In the 1980’s, the speed and capacity of computers increased dramatically, and, with

more and more texts being produced in computerized form, it became possible to create

corpora much larger than the Brown and LOB, containing millions of words. The

availability of language samples of this magnitude opened up the possibility of gathering

meaningful statistics about language patterns that could be used to drive language

processing software such as syntactic parsers, which sparked renewed interest in corpus

compilation within the computational linguistics community. Parallel corpora, which

contain the same text in two or more languages, also began to appear; the best known of

these is the Canadian Hansard corpus of Parliamentary debates in English and French.

Corpus creation still involved considerable work, even when texts could be acquired from

other sources in electronic form. For example, many texts existed as typesetter’s tapes

obtained from publishers, and substantial processing was required to remove or translate

typesetter codes.

The “golden era” of linguistic corpora began in 1990 and continues to this day. Enormous

corpora of both text and speech have been and continue to be compiled, many by

government-funded projects in Europe, the U.S., and Japan. In addition to mono-lingual

corpora, several multi-lingual parallel corpora covering multiple languages have also

been created. A side effect of the growth in the availability and use of corpora in the

1990’s was the development of automatic techniques for annotating language data with

information about its linguistic properties. Algorithms for assigning part of speech tags to

words in a corpus and aligning words and sentences in parallel text (i.e., associating each

word or sentence with its translation in the parallel version) were developed in the 1990’s

that achieve 95-98% accuracy. Automatic means to identify syntactic configurations such

as noun phrases, and proper names, dates, etc., were also developed.

There now exist numerous corpora, many of which are available through the Linguistic

Data Consortium (LDC) (http://www.ldc.upenn.edu) in the U.S. and the European

Language Resources Association (ELRA) (http://www.elra.org) in Europe, both of which

were founded in the mid-nineties to serve as repositories and distributors of corpora and

other language resources such as lexicons. However, because of the cost and difficulty of

obtaining some types of texts (e.g., fiction), existing corpora vary considerably in their

composition; very few efforts have been made to compile language samples that are

“balanced” in their representation of different genres. Notable exceptions (apart from the

early Brown and LOB) are the British National Corpus (BNC)

(http://www.hcu.ox.ac.uk/BNC/) and the American National Corpus (ANC)

(http://www.AmericanNationalCorpus.org), as well as (to some extent) the corpora for

several Western European languages produced by the PAROLE project. In fact the

greatest number of existing text corpora are composed of readily available materials such

as newspaper data, technical manuals, government documents, and, more recently,

materials drawn from the World Wide Web. Speech data, whose acquisition is in most

instances necessarily controlled, are more often representative of a specific dialect or

range of dialects.

Many corpora are available for research purposes by signing a license and paying a small

reproduction fee. Other corpora are available only by paying a (sometimes substantial)

fee; this is the case, for instance, for many of the holdings of the LDC, making them

virtually inaccessible to humanists.

Preparation of linguistic corpora

The first phase of corpus creation is data capture, which involves rendering the text in

electronic form, either by hand or via OCR, acquisition of word processor or publishing

software output, typesetter tapes, PDF files, etc. Manual entry is time-consuming and

costly, and therefore unsuitable for the creation of very large corpora. OCR output can be

similarly costly if it requires substantial post-processing to validate the data. Data

acquired in electronic form from other sources will almost invariably contain formatting

codes and other information that must be discarded or translated to a representation that is

processable for linguistic analysis.

Representation formats and surrounding issues

At this time, the most common representation format for linguistic corpora is XML.

Several existing corpora are tagged using the EAGLES XML Corpus Encoding Standard

(XCES) (Ide, 1998), a TEI-compliant XML application designed specifically for

linguistic corpora and their annotations. The XCES introduced the notion of stand-off

annotation, which requires that annotations are encoded in documents separate from the

primary data and linked to it. One of the primary motivations for this approach is to avoid

the difficulties of overlapping hierarchies, which are common when annotating diverse

linguistic features, as well as the unwieldy documents that can be produced when

multiple annotations are associated with a single document. The stand-off approach also

allows for annotation of the same feature (e.g., part of speech) using alternative schemes,

as well as associating annotations with other annotations rather than directly to the data.

Finally, it supports two basic notions about text and annotations outlined in Leech (1993):

it should be possible to remove the annotation from an annotated corpus in order to revert

to the raw corpus; and. conversely, it should be possible to extract the annotations by

themselves from the text.

The use of stand-off annotation is now widely accepted as the norm among corpus and

corpus-handling software developers; however, because mechanisms for inter-document

linking have only recently been developed within the XML framework, many existing

corpora include annotations in the same document as the text.

The use of the stand-off model dictates that a distinction is made between the primary

data (i.e., the text without additional linguistic information) and its annotations, in

particular, what should and should not be marked in the former. The XCES identifies two

types of information that may be encoded in the primary data:

1. Gross structure: universal text elements down to the level of paragraph, which is

the smallest unit that can be identified language-independently; for example,

• structural units of text, such as volume, chapter, etc., down to the level of

paragraph; also footnotes, titles, headings, tables, figures, etc.;

• features of typography and layout, for previously printed texts: e.g., list item

markers;

• non-textual information (graphics, etc.).

2. Segmental structure: elements appearing at the sub-paragraph level which are

usually signalled (sometimes ambiguously) by typography in the text and which

are language dependent; for example.

• orthographic sentences, quotations;

• orthographic words;

• abbreviations, names, dates, highlighted words.

Annotations (see next section) are linked to the primary data using XML conventions

(XLink, Xpointer).

Speech data, especially speech signals, is often treated as “read-only”, and therefore the

primary data contains no XML markup to which annotations may be linked. In this case,

stand-off documents identify the start and end points (typically using byte offsets) of the

structures listed above, and annotations are linked indirectly to the primary data by

referencing the structures in these documents. The annotation graphs representation

format used in the ATLAS project, which is intended primarily to handle speech data,

relies entirely on this approach to link annotations to data, with no option for referencing

XML-tagged elements.

Identification of segmental structures

Markup identifying the boundaries of gross structures may be automatically generated

from original formatting information. However, in most cases the original formatting is

presentational rather than descriptive; for example, titles may be identifiable because

they are in bold, and therefore transduction to a descriptive XML representation may not

be straightforward. This is especially true for sub-paragraph elements that are in italic or

bold font; it is usually impossible to automatically tag such elements as emphasis, foreign

word, etc.

Creation of linguistic corpora almost always demands that sub-paragraph structures such

as sentences and words, as well as names, dates, abbreviations, etc., are identified.

Numerous programs have been developed to perform sentence “splitting” and word

tokenization, many of which are freely available (see, for example, the tools listed in the

Natural Language Software Registry (http://www.dfki.de/lt/registry/) or the SIL Software

Catalog (http://www.sil.org). These functions are also embedded in more general corpus

development tools such as GATE (Cunningham, 2002). Sentence splitting and

tokenization are highly language-dependent and therefore require specific information

(e.g., abbreviations for sentence splitting, clitics and punctuation conventions for

tokenization) for the language being processed; in some cases, language-specific software

is developed, while in others a general processing engine is fed the language-specific

information as data and can thus handle multiple languages. Languages without word-

boundary markers, such as Chinese and Japanese, and continuous speech represented by

phoneme sequences, require an entirely different approach to segmentation, the most

common of which is a dynamic programming algorithm to compute the most likely

boundaries from a weighted transition graph. This of course demands that the

probabilities of possible symbol or phoneme sequences are available in order to create the

weighted graph.

Within the computational linguistics community, software to identify so-called “named

entities” (proper names designating people, places, organizations, events, documents,

etc.), as well as dates and other time expressions, has been developed, and many of these

tools are freely available for research. However, most of these tools have been developed

using existing corpora which consist of newspapers and government reports and are

unlikely to perform well on the kinds of data that interest humanists, such as literary

works, historical documents, etc. This is just one example of the broader situation in the

development of corpus-handling tools: the available data, which are often highly skewed

for genre, drive tool and algorithm development, and therefore applicability to more

generalized corpora is often limited.

Corpus annotation

For computational linguistics research, which has driven the bulk of corpus creation

efforts over the past decade, corpora are typically annotated with various kinds of

linguistic information. The following sections outline the major annotation types.

Morpho-syntactic annotation

By far the most common corpus annotation is morpho-syntactic annotation (part-of-

speech tagging), primarily because several highly accurate automatic taggers have been

developed over the past 15 years. Part of speech tagging is a disambiguation task: for

words that have more than one possible part of speech, it is necessary to determine which

one, given the context, is correct. Although in English, close to 90% of the words in the

language have only one part of speech, in actual use (e.g., in a corpus), part of speech for

as many as 40% of the word instances may be ambiguous (DeRose, 1988), due in large

part to ambiguity for a handful of high-frequency words such as “that”, which can be

either a determiner (“that girl”) or a complementizer (“He heard that you came home”).

Beyond this, the most common ambiguity in English is between verb and noun, e.g.,

“hide”, “dog”, “brand”, “felt”, etc.

Taggers fall into two general classes: rule-based (e.g. ENGTWOL Voutlianen, 1995),

which use manually generated rules to assign part of speech, and stochastic, which rely

on probabilities for n-grams—sequences of n (usually, 2 or 3) tags that are known to

occur in real data. Stochastic taggers learn these probabilities by being “trained” on

previously tagged data that has been hand-validated for correctness. A third type of

tagger, often called the “Brill tagger” after its developer (Brill, 1995), uses a hybrid

approach which learns its tagging rules from a previously tagged training corpus.

Obviously, the more accurately tagged data a tagger can use for training, the more likely

its probabilities will be correct. This fact has led to the creation of corpora that have been

hand-annotated (or in which automatically produced tags have been hand-validated)

specifically intended for training, in order to enable automatic taggers to produce even

more tagged corpora.

The tags generated by a part of speech tagger provide more information than simple word

class (noun, verb, adjective, etc.). Various levels of morpho-syntactic information can be

represented; the more detailed the information, the bigger the tagset, and the bigger the

tagset, the less accurate an automatic tagger will be. For this reason, tagsets including 50-

100 tags that collapse or eliminate detailed morpho-syntactic information—such as the

information in the morpho-syntactic specifications for Western and eastern European

languages produced by the EAGLES project

(http://www.ilc.pi.cnr.it/EAGLES/home.html) -- are most typically used in automatic

tagging.

There are several tagsets in common use for English, most of which evolved from the 87

tags used in the Brown corpus (http://www.hit.uib.no/icame/brown/bcm.html). Probably

the most widely used is the 45-tag set of the Penn Treebank project (Marcus, et al.,

1993), of which only 36 are actual morpho-syntactic categories (the rest are for

punctuation, list markers, etc.). The Penn tagset is a variant of the Brown tagset that

eliminates information retrievable from the form of the lexical item. It therefore includes

only one tag for different forms of the verbs “be”, “have”, and “do”, whereas the Brown

(and other common tagsets for English) provide a different tag for each of these forms.

Another well-known tagset for English is the 61-tag C5 tagset of the CLAWS

(Constituent Likelihood Automatic Word-tagging System) tagger, developed at the

University of Lancaster (Garside, 1997) and used to tag the British National Corpus.

Part of speech taggers rely on lexicons that provide all of the possible part of speech

assignments for a lexical item found in the input, from which it must choose the most

probable given the immediate context. The information in the lexicons must therefore

match—or at least, be mappable to—the tagset used by the tagger. It is an unfortunate

fact that it is often extremely difficult and sometimes impossible to map one tagset to

another, which has resulted in much re-creation of lexical information to suit the needs of

a particular tagger.

Many lexicons include lemmas (root forms), and morpho-syntactic annotation may

produce lemmas as well as part of speech tags in their output. The presence of lemmas in

an annotated text enables extraction of all orthographic forms associated with a given

lemma (e.g., “do”, “does”, “doing, “did” for the lemma “do”). However, although

relatively easily produced, many existing corpora do not include lemmas; notable

exceptions are the Journal of the Commission corpus, the Orwell multi-lingual corpus,

and the SUSANNE corpus.

Parallel alignment

In addition to algorithms for morpho-syntactic annotation, reliable algorithms for

alignment of parallel texts—i.e., texts for which there exist translations in two or more

languages—have been developed. Probability information about correspondences

between words, phrases, and other structures derived from aligned corpora are used to

choose from multiple possible translations that may be generated by a machine

translation system. Parallel corpora have also been used to automatically generate

bilingual dictionaries (e.g., Dagan and Church, 1997) and, more recently, as a means to

achieve automatic sense tagging of corpora (e.g., Ide, et al., 2001).

Two types of parallel alignment are common: sentence alignment and word alignment.

Sentence alignment is by far the easier and more accurate of the two, the major problem

being to determine cases in which one-to-many or partial mappings exist. Many sentence

and word-aligned corpora exist, the vast majority of which include only two languages.

Probably the best known is the English-French Hansard Corpus of Canadian

Parliamentary debates, which has served as the basis for numerous translation studies.

Multi-lingual parallel corpora are much more rare; the difficulty is not in the alignment

itself, but rather in the availability of texts in multiple translations (in particular, texts that

are not bound by copyright or other restrictions). Existing multi-lingual aligned corpora

include the United Nations Parallel Text Corpus (English, French, Spanish), the Journal

of the Commission (JOC) corpus (English, French, German, Italian, Spanish), the Orwell

1984 Corpus (Bulgarian, Czech, Estonian, Hungarian, Romanian, Slovene, English,

Latvian, Lithuanian, Serbo-Croat), Plato's Republic (Bulgarian, Chinese, Czech, English,

German, Latvian, Polish, Romanian, Slovak, Slovene) and the Bible (Chinese, Danish,

English, Finnish, French, Greek, Indonesian, Latin, Spanish, Swahili, Swedish,

Vietnamese).

Syntactic annotation

There are two main types of syntactic annotation in linguistic corpora: noun phrase (NP)

bracketing or “chunking”, and the creation of “treebanks” that include fuller syntactic

analysis. Syntactically annotated corpora serve various statistics-based applications, most

notably, by providing probabilities to drive syntactic parsers, and have been also used to

derive context-free and unification-based grammars (Charniak, 1996; van Genabith et al.,

1999). Syntactically annotated corpora also provide theoretical linguists with data to

support studies of language use.

The best known and most widely used treebank is the Penn Treebank for English (Marcus

et al., 1994). Ongoing projects exist for the development of treebanks for other

languages, including German (the NEGRA corpus, Brants, et al., 2003; Verbmobil,

Stegmann et al., 1998), Czech (The Prague Dependency Treebank, Hajic, 1998), French

(Abeillé, 2000), and Chinese (Xue et al., 2002). The Penn Treebank and the Chinese

Treebank, both created at the University of Pennsylvania, use LISP-like list structures to

specify constituency relations and provide syntactic category labels for constituents, as

shown below:

((S (NP-SBJ-1 Jones)

(VP followed)

(NP him)

(PP-DIR into

(NP the front room))

,

(S-ADV (NP-SBJ *-1)

(VP closing

(NP the door)

(PP behind

(NP him)))))

.))

Although they differ in the labels and in some cases the function of various nodes in the

tree, many treebank annotation schemes provide a similar constituency-based

representation of relations among syntactic components. In contrast, dependency schemes

do not provide a hierarchical constituency analysis, but rather specify grammatical

relations among elements explicitly; for example, the sentence “Paul intends to leave

IBM” could be represented as follows:

subj(intend,Paul,_)

xcomp(intend,leave,to)

subj(leave,Paul)

dobj(leave,IBM,_)

Here, the predicate is the relation type, the first argument is the head, the second the

dependent, and additional arguments may provide category-specific information (e.g.,

introducer for prepositional phrases). Finally, so-called “hybrid systems” combine

constituency analysis and functional dependencies, usually producing a shallow

constituent parse that brackets major phrase types and identifies the dependencies

between heads of constituents (e.g. the NEGRA Corpus).

Although most modern treebanks utilize an SGML or XML encoding rather than list

structures, syntactic annotation is invariably interspersed with the text itself. This makes

it difficult or impossible to add other kinds of annotations to the data, or to provide

alternative syntactic annotations. As noted earlier, the use of stand-off annotations is

increasingly encouraged. A stand-off scheme for syntactic annotations, which also serves

as a “pivot” format for representing different types of syntactic annotation with a

common scheme, has been developed by Ide and Romary (2001).

Semantic annotation

Semantic annotation can be taken to mean any kind of annotation that adds information

about the meaning of elements in a text. Some annotations that may be considered to

provide semantic information—for example, “case role” information such as agent,

instrument, etc.--are often included in syntactic annotations. Another type of semantic

annotation common in literary analysis (especially in the 1960’s and 70’s) marks words

or phrases in a text as representative of a particular theme or concept. At present, the

most common type of semantic annotation is ”sense tagging”: the association of lexical

items in a text with a particular sense or definition, usually drawn from an existing sense

inventory provided in a dictionary or on-line lexicon such as WordNet (Miller, et al.,

1990).

The major difficulty in sense annotation is to determine an appropriate set of senses.

Simply examining the differences in sense distinctions made from one dictionary to

another demonstrates the difficulty of this task. To solve the problem, some attempts

have been made to identify the kinds of sense distinctions that are useful for automatic

language processing: for example, tasks such as information retrieval may require only

very coarse-grained sense distinctions—the difference between “bank” as a financial

institution vs. a river bank—whereas others, and in particular machine translation, require

finer distinctions—say, between “bank” as a financial institution and as a building.

However, by far the most common source of sense tags used for semantic annotation is

WordNet, an on-line dictionary that in addition to providing sense lists groups words into

“synsets” of synonomous words. WordNet has been updated several times, and as a

result, its sense list may differ for a particular word depending on the WordNet version

used for the tagging.

It is widely acknowledged that the sense distinctions provided in WordNet are far from

optimal. However, this resource, which is among the most heavily used in natural

language processing research over the past decade, will likely continue to serve as a basis

for sense tagging for at least the foreseeable future, if for no other reason than that it

continues to be the only freely available, machine tractable lexicon providing extensive

coverage of English. The EuroWordNet project has produced WordNets for most

Western European languages linked to WordNet 1.5 (the current version of the English

WordNet is 2.0), and WordNets for other languages (e.g., Balkan languages) are under

development, which will likely extend the research community’s reliance on its sense

inventory. In any case, no clearly superior alternative source of sense distinctions has

been proposed.

Because sense-tagging requires hand annotation, and because human annotators will

often disagree on sense assignments even given a pre-defined sense inventory, very few

sense-tagged corpora exist. Examples include the Semantic Concordance Corpus

(SemCor; Miller, et al., 1993), produced by the WordNet project, which assigns WordNet

sense tags for all nouns, verbs, and adjectives in a 250,000 word corpus drawn primarily

from the Brown Corpus; the DSO Corpus, containing sense-tags for 121 nouns and 70

verbs in about 192,800 sentences taken from the Brown Corpus and the Wall Street

Journal; and Hector, containing about 200,000 tagged instances of 300 words in a corpus

of British English.

Automatic means to sense-tag data have been sought ever since machine-readable texts

became available. This area of research, called “word sense disambiguation”, remains to

this day one of the more difficult problems in language processing. Although rule-based

approaches have been developed, the most common approaches to word sense

disambiguation over the past decade are statistics-based, relying on the frequency with

which lexical items (or categories of lexical items) in the context under consideration

have been found in the context of a word in a given sense. Some recent research (Ide,

1999; Ide et al., 2002; Doab and Resnik, 2002) has explored the use of information

gleaned from parallel translations to make sense distinctions. For a comprehensive

overview of approaches to word sense disambiguation see Ide and Véronis, 1998.

Discourse-level annotation

There are three main types of annotation at the level of discourse: topic identification, co-

reference annotation, and discourse structure.

Topic identification (also called “topic detection”) annotates texts with information about

the events or activities described in the text. Automatic means for topic detection over

streams of data such as broadcast news and newswires is under development, primarily in

the context of the DARPA-sponsored Topic Detection Task. A sub-task of this kind of

annotation is detection of boundaries between stories/texts, which may also be included

in the annotation.

Co-reference annotation links referring objects (e.g., pronouns, definite noun phrases) to

prior elements in a discourse to which they refer. This type of annotation is invariably

performed manually, since reliable software to identify co-referents is not available. For

example, the MUC 7 (Message Understanding Conference) corpus contains hand-

generated co-reference links (Hirschman and Chinchor, 1997) marked in a non-stand-off

SGML format similar to the following:

<COREF ID="E5">This appliance</COREF> is supplied with

<COREF ID="E6">a fitted three pin mains plug</COREF>.

<COREF ID="E7">A 3 amp fuse</COREF>

is fitted in <COREF ID="E8" TYPE="IDENT" REF="E6">the plug</COREF>.

Should <COREF ID="E9" TYPE="IDENT" REF="E7">the fuse</COREF>

need to be replaced…

Discourse structure annotation identifies multi-level hierarchies of discourse segments

and the relations between them, building upon low-level analysis of clauses, phrases, or

sentences. There are several theoretical approaches to the analysis of discourse structure,

differing in terms of the span of text (clause, sentence, etc.) that is considered to be the

atomic unit of analysis, and in the relations defined to exist between units and higher-

level structures built from them. Most common approaches are based on “focus spaces”

(Grosz and Sidner, 1986) or Rhetorical Structure Theory (Mann and Thompson, 1987).

To date, annotation of discourse structure is almost always accomplished by hand,

although some software to perform discourse segmentation has been developed (e.g.,

Marcu, 1996).

Annotation of speech and spoken data

The most common type of speech annotation is an orthographic transcription that is time-

aligned to an audio or video recording. Often, annotation demarcating the boundaries of

speech “turns” and individual “utterances” is included. Examples of speech corpora in

this type of format include the Child Language Data Exchange System (CHILDES)

corpus (MacWhinney, 1995), the TIMIT corpus of read speech (Lamel, et al., 1996), and

the LACITO Linguistic Data Archiving Project (Jacobsen and Michailovsky, 2000).

The orthographic transcription of speech data can be annotated with any of the kinds of

linguistic information outlined in previous sections (e.g., part of speech, syntax, co-

reference, etc.), although this is not common. In addition, speech data may be annotated

with phonetic segmentation and labeling (phonetic transcription), prosodic phrasing and

intonation, disfluencies, and, in the case of video, gesture.

Most types of annotation particular to speech data are time-consuming to produce

because they must be generated by hand, and because the annotators must be skilled in

the recognition and transcription of speech sounds. Furthermore, speech annotation is

problematic: for example, phonetic transcription works on the assumption that the speech

signal can be divided into single, clearly demarcated sounds, but these demarcations are

often rather unclear. Prosodic annotation is even more subjective, since the decision

about the exact nature of a tone movement often varies from annotator to annotator. The

kinds of phenomena that are noted include onset, rising tone, falling tone, rising/falling

tone, level tone, pause, overlap, etc. The notations for prosodic annotation vary widely

and they are not typically rendered in a standard format such as XML. As result, the few

existing corpora annotated for prosody are widely inconsistent in their format. One of the

best known such corpora is the London-Lund Corpus of Spoken English (Svartvik, 1990).

Corpus annotation tools

Over the past decade, several projects have created tools to facilitate the annotation of

linguistic corpora. Most are based on a common architectural model introduced in the

MULTEXT project, which views the annotation process as a chain of smaller, individual

processes that incrementally add annotations to the data. A similar model was developed

in the TIPSTER project.

Among the existing annotation tools for language data are LT XML (University of

Edinburgh), which directly implements the MULTEXT model and is especially geared

toward XML-encoded resources; GATE (University of Sheffield), based on the TIPSTER

model and including tools for tokenization, sentence splitting, named entity recognition,

part of speech tagging, as well as corpus and annotation editing tools. The Multilevel

Annotation, Tools Engineering (MATE) project provides an annotation tool suite

designed especially for spoken dialogue corpora at multiple levels, focusing on prosody,

(morpho-) syntax, co-reference, dialogue acts, and communicative difficulties, as well as

inter-level interaction. ATLAS (Architecture and Tools for Linguistic Analysis Systems)

is a joint initiative of the US National Institute for Standards and Technology (NIST),

MITRE, and the LDC to build a general purpose annotation architecture and a data

interchange format. The starting point in ATLAS is the annotation graph model, with

some significant generalizations.

Currently, a sub-committee of the International Standards Organization (ISO)—ISO

TC37 SC4--is developing a generalized model for linguistic annotations and processing

tools, based on input from developers of the annotation tool suites mentioned above as

well as from annotation scheme designers. The goal is to provide a common “pivot”

format that instantiates a generalized data model of linguistic resources and their

annotations to and from which existing formats—provided they are consistent with the

model—can be mapped in order to enable seamless interchange. At present, data

annotated within one project, using XML, annotation graphs, or a proprietary annotation

scheme is typically difficult to import for manipulation using another tool suite, either for

further annotation or analysis. The developing ISO model is intended to allow annotators

to use any of a variety of schemes to represent their data and annotations, and map it to a

common format for interchange. Users of other schemes would then import the data in

the common format to their own schemes. This means, for example, that a corpus marked

up using TEI or XCES conventions can be mapped to the common format, and from that,

mapped again to, say, an annotation graph representation that will enable manipulation of

the data by tools implementing that model, without information loss.

The future of corpus annotation

Recent developments in the XML world, primarily in the scope of work within the World

Wide Web Consortium (W3C), have focused attention on the potential to build a

Semantic Web. This possibility has interesting ramifications for both corpus annotation

and analysis, in two (related) ways. First, the underlying technology of the Semantic Web

enables definition of the kinds of relationships (links) one resource—where a “resource”

can be any fragment of a document or the document as a whole--may have with another.

For example, a “word” may have a link labeled “part of speech” with another resource

that represents (possibly, as a simple string) “noun” or “noun singular masculine”.

Because annotation is, at the base, the specification of relationships between information

in a corpus and linguistic information that describes it, the development of technologies

such as the Resource Description Framework (RDF) can have significant impact on the

way annotations are associated with primary language resources in the future.

A second activity within the Semantic Web community that has ramifications for both

annotation and analysis of linguistic data is the development of technologies to support

the specification of and access to ontological information. Ontologies, which provide a

priori information about relationships among categories of data, enable the possibility to

apply inferencing processes that can yield information that is not explicit in the data

itself. For example, to properly parse a sentence such as “I ate a fish with a fork”—that is,

to attach the prepositional phrase “with a fork” to the verb “eat” and not as a modifier of

“fish”--we can check an ontology that specifies that “fork” IS-A sub-class of

“instrument”, and that “eat” has a USES-A relation with things of type “instrument”.

More generally, we may be able to identify the topic of a given document by consulting

an ontology. For example, the ontology may specify that the word “bank” can represent a

sub-class of “financial institution” or “geological formation”; if the document contains

several other words that are also related to “financial institution” such as “money”,

“account”, etc., we may conclude that the document has financial institutions as one of its

topics (and, as a side effect, that the word “bank” is being used in that sense in this

document).

In order to be able to exploit ontological information, it must be created and stored in a

universally accessible form. The W3C group developing the Ontological Web Language

(OWL) is providing a standard representation format. It is now up to the computational

linguistics community to instantiate the relevant ontological information and use it to

annotate and analyze language data. Development and use of ontologies has in fact been

a part of the field for several years—in particular, the ontology in the WordNet lexicon

has been widely used for annotation and analysis of language data. Semantic Web

technologies will enable development of common and universally accessible ontological

information.

Analysis of linguistic corpora

There are two major uses for linguistic corpora: statistics-gathering to support natural

language processing; and language analysis to support language learning and dictionary

creation.

Statistics-gathering

A corpus provides a bank of samples that enable the development of numerical language

models, and thus the use of corpora goes hand-in-hand with empirical methods. In the

late 1980s, the increased availability of large amounts of electronic text enabled, for the

first time, the full-scale use of data-driven methods to attack generic problems in

computational linguistics, such as part-of-speech identification, prepositional phrase

attachment, parallel text alignment, word sense disambiguation, etc. The success in

treating at least some of these problems with statistical methods led to their application to

others, and by the mid-1990s, statistical methods had become a staple of computational

linguistics work.

The key element for many statistics-based approaches to language processing (in

particular, so-called “supervised” methods) is the availability of large, annotated corpora,

upon which annotation algorithms are trained to identify common patterns and create

transformations or rules for them. The stochastic part of speech taggers described above

are probably the most well-known application relying on previously annotated corpora,

but similar approaches are also used for word sense disambiguation, probabilistic parsing,

and speech recognition. In word sense disambiguation, for example, statistics are

gathered reflecting the degree to which other words are likely to appear in the context of

some previously sense-tagged word in a corpus. These statistics are then used to

disambiguate occurrences of that word in untagged corpora, by computing the overlap

between the unseen context and the context in which the word was seen in a known

sense.

Speech recognition is, in fact, the area of computational linguistics in which corpora were

first put into large-scale use in support of language processing, beginning with the

utilization of Hidden Markov Models (HMMs) in the 1970s. This paradigm required data

to statistically train an acoustic model to capture typical sound sequences and a language

model to capture typical word sequences, and produced results that were far more

accurate and robust than the traditional methods. It was not until the late 1980’s that the

statistical approach was applied to other areas, one of the first of which was machine

translation. Following the same approach as the speech recognition systems, researchers

automatically trained a French-English correspondence model (the Translation Model) on

3 million sentences of parallel French and English from the Canadian Parliamentary

records, and also trained a Language Model for English production from Wall Street

Journal data. To translate, the former model was used to replace French words or phrases

by the most likely English equivalents, and then the latter model ordered the English

words and phrases into the most likely sequences to form output sentences.

Probabilistic parsing is one of the more recent applications of statistical methods to

language processing tasks. Again, large bodies of data previously annotated and validated

for syntactic structure are required in order to provide statistics concerning the probability

that a given syntactic construction is the correct one in its context. Syntactic structure can

be very ambiguous; traditional parsers often produced numerous alternative structural

analyses for an input sentence. A probabilistic parser uses previously gathered statistics

to choose the most probable interpretation.

Issues for corpus-based statistics gathering

In order to be representative of any language as a whole, it is necessary that a corpus

include samples across a variety of texts that reflect the range of syntactic and semantic

phenomena across that language. This demands, first of all, that the data be adequately

large in order to avoid the problem of data sparseness that plagues many statistical

approaches. For example, for tasks such as word sense disambiguation, data must be

extensive enough to ensure that all senses of a polysemous word are not only represented,

but also represented frequently enough so that meaningful statistics can be compiled.

Although it has been extensively used for natural language processing work, the million

words of a corpus such as the Brown Corpus are not sufficient for today's large-scale

applications: many word senses are not represented; many syntactic structures occur too

infrequently to be significant, and the corpus is far too small to be used for computing the

bi-gram and tri-gram probabilities that are necessary for training language models for

speech recognition.

Unfortunately, in the main, the large corpora freely available for research consist of texts

that can be easily acquired and are available for re-distribution without undue problems

of copyright, etc. Because of this, corpora used for statistics-gathering for language

processing are vastly over-representative of certain genres, in particular newspaper

samples, which comprise the greatest percentage of texts currently available from, for

example, the LDC, and which also dominate the training data available for speech

recognition purposes. Other available corpora typically consist of technical reports,

transcriptions of parliamentary and other proceedings, short telephone conversations, and

the like. The upshot of this is that corpus-based natural language processing has relied

heavily on language samples representative of usage in a handful of limited and

linguistically specialized domains. This can lead to drastically skewed results: for

example, in newspaper data, there is a disproportionate number of complex NP

complements for some verbs, which appear in sentences typical of newspaper style, such

as "The price rose two percent to 102 dollars per share from 100 dollars per share".

Similar problems arise in work on word sense disambiguation: it has been noted that for

some typical test words such as "line", certain senses (for example, the common sense of

"line" as in the sentence, "He really handed her a line") are absent entirely from resources

such as the Wall Street Journal.

The problem of balance is acute in speech recognition. Speech recognition systems are

notoriously dependent on the characteristics of their training corpora. Corpora large

enough to train the trigram language models of modern speech recognizers (many tens of

millions of words) are invariably composed of written rather than spoken texts. But the

differences between written and spoken language are even more severe than the

differences balanced corpora like the Brown and newspaper corpora like the Wall Street

Journal. Therefore, whenever a state-of-the-art speech recognition research effort moves

to a new domain, a new large training corpus of speech must be collected, transcribed at

the word level, and the transcription must be aligned to the speech.

Language analysis

The gathering of authentic language data from corpora enables a description of language

that starts from the evidence rather than from imposing some theoretical model. Because

speakers and writers produce language with real communicative goals, corpora of native

speaker texts provide, in principle, samples of genuine language. For this reason, one of

the oldest uses of corpora is for dictionary–making, or lexicography, and in particular,

lexicography with the goal of producing so-called “learners’ dictionaries” designed for

those learning a new language.

The COBUILD corpus was compiled starting in the early 1980’s, at which time it

included about 7 million words. This corpus was used to create the Collins COBUILD

English Dictionary, the first dictionary relying fully on corpora for its creation. Following

this lead, over the course of the next decade most British dictionary publishers began to

use corpora as the primary data-source for their dictionaries, although interestingly,

American dictionary publishers are only now beginning to rely on corpora to guide

lexicography.

The basic lexicographic tool for analyzing a corpus is a concordancer, a program that

displays occurrences of a word in the middle of a line of context from the corpus.

However, the huge increase in available data has led to a situation where lexicographers

are presented with hundreds or even thousands of concordance lines for a single word. As

a result, lexicographers have begun to rely on techniques derived from computational

linguistics to summarize concordance data. The most common is the “Mutual

Information” (MI) score, a statistical measure that shows how closely one word is

associated with others based on the regularity with which they co-occur in context. For

example, in English, the mutual information score for the words "strong" and "tea" is

much higher than the score for "powerful" and "tea", despite the fact that “strong” and

“powerful” have similar meanings. This kind of information is invaluable for

lexicographers, especially for the creation of language learners’ dictionaries that must

provide this kind of distinction.

Most recently, dictionary-makers have teamed with researchers in the field of

computational linguistics to glean even more precise information from corpus data. For

example, MI scores can show that “strong” collocates with northerly, showings, believer,

currents, supporter and odor, while “powerful” collocates with words such as tool,

minority, neighbor, symbol, figure, weapon and post; but the MI score does not indicate

the kinds of grammatical relations that exist among a word and its collocates.

Supplemental grammatical information can provide even more precise understanding of

word usage. For example, the word “tea” collocates with words like spoon, milk, and

sugar when these words appear as the object of the preposition “with”; with mug, saucer,

and cup when they are objects of the preposition “into”; with coffee, toast, sugar, etc.

when joined by “and” or “or”; and “tea” is most commonly the object of verbs such as

drink, sip, pour, finish, and make.

To gather this information, relatively sophisticated language processing software is

required that can annotate the data for one or more of the types of information outlined in

the previous sections. The need for more informative results to serve the needs of

lexicography has therefore led to increased collaboration with computational linguists.

We see, as a result, two groups of researchers that had previously worked independently

coming together to tackle common problems.

In general, researchers in the humanities and researchers in computational linguistics

have not collaborated, despite their common problems and goals. With the advent of the

World Wide Web, this should change soon. Humanists have increased access to

information about work in computational linguistics as well as to tools and resources

developed by that community. Computational linguists, on the other hand, are likely to

face new language processing challenges due to the need to handle a greater variety of

web-accessible materials, including literary works, historical documents, and the like.

The eventual collaboration of these two groups should lead, in the end, to vastly

increased capabilities for both.

References for Further Reading

Abeillé, A, Clément, L., Kinyon, A. (2000). Building a treebank for French. Proceedings

of the Second International Conference on Language Resources and Evaluation, 87-94.

Brants, T., Skut, W., Uszkoreit, H. (in press). Syntactic annotation of a German

newspaper corpus. In A. Abeillé (Ed.), Building and Using Syntactically Annotated

Corpora. Dordrecht: Kluwer Academic Publishers.

Brill, E. (1995). Transformation-based error-driven learning and natural language

processing: A case study in part of speech tagging. Computational Linguistics, 21, 543-

565.

Charniak, E. (1996). Tree-bank grammars. AAAI-96: Proceedings of the Thirteenth

National Conference on Artificial Intelligence, 598-603.

Cunningham, H. (2002). GATE, a general architecture for text engineering. Computers

and the Humanities, 36, 223-254.

Dagan,I. Church, K. (1997). Termight: Coordinating humans and machines in bilingual

terminology translation. MachineTranslation, 12, 89-107.

DeRose, S. (1988). Grammatical category disambiguation by statistical optimization.

Computational Linguistics, 14. 31-39.

Doab, M., Resnik,, P. (2002). An unsupervised method for word sense tagging using

parallel corpora. Proceedings of the 40th Annual Meeting of the Association for

Computational Linguistics, 255-261.

Garside, R., Smith, N. (1997). A hybrid grammatical tagger: CLAWS4. In R. Garside, G.

Leech, & A. McEnery (Eds). Corpus Annotation: Linguistic Information from Computer

Text Corpora (pp.102-121). London: Longman.

Grosz, B., Sidner, C. (1986). Attention, intention and the structure of discourse.

Computational Linguistics, 12, 175-204.

Hajic, J. (1998). Building a syntactically annotated corpus: The Prague Dependency

Treebank. In E. Hajicova (Ed.), Issues of Valency and Meaning. Studies in Honour of

Jarmila Panevova. Prague: Charles University Press.

Hirschman L., Chinchor, N. (1997). MUC-7 Co-reference task definition. Version 3.0,

http://www.itl.nist.gov/iaui/894.02/related_projects/muc/proceedings/co_task.html.

Ide, N. Erjavec, T., Tufis, D. (2001). Automatic sense tagging using parallel corpora.

Proceedings of the Sixth Natural Language Processing Pacific Rim Symposium, 83-89.

Ide, N., Bonhomme, P., Romary, L. (2000). XCES: An XML-based standard for

linguistic corpora.. Proceedings of the Second Language Resources and Evaluation

Conference, 825-30.

Ide, N., Romary, L. (2001). A common framework for syntactic annotation. Proceedings

of ACL'2001, 298-305.

Ide, N., Veronis, J. (1998) Introduction to the special issue on word sense

disambiguation: The state of the art. Computational Linguistics, 24, 1-40.

Jacobson, M., Michailovsky B. (2000). A Linguistic archive on the web.

IRCS/ISLE/Talkbank Linguistic Exploration Workshop: Web-Based Language

Documentation and Description.

http://www.ldc.upenn.edu/exploration/expl2000/papers/michailovsky/index.htm)

Johansson, S., Leech, G., Goodluck, H. (1978). Manual of information to accompany the

Lancaster-Oslo/Bergen Corpus of British English, for use with digital computers.

Department of English, University of Oslo.

Kucera H., and Francis, W.N. (1967). Computational Analysis of Present-Day American

English. Providence, R.I., Brown University Press.

Lamel, L.F., Garafolo, J., Fiscus, J,. Fisher, W., Pallett. D. S. (1990) TIMIT: The Darpa

acoustic-phonetic speech corpus. Technical Report NISTIR 4930, National Institute of

Standards and Technology.

Leech, G. (1993) Corpus annotation schemes. Literary and Linguistic Computing 8, 275-

81.

MacWhinney, B. (1995). The CHILDES Project (2nd ed.). Mahwah, NJ: Lawrence

Erlbaum.

Mann, W.C., Thompson S.A. (1988). Rhetorical structure theory: A theory of text

organization. Text, 8, 243-281.

Marcu. D. (1996). Building up rhetorical structure trees. Proceedings of the Thirteenth

National Conference on Artificial Intelligence, 2, 1069-1074.

Marcus, M., Santorini, B., Marcinkiewicz., M. A. (1993). Building a large annotated

corpus of English: The Penn Treebank. Computational Linguistics, 19(2).

Miller, G., Leacock, C., Randee, T. and Bunker, R. (1993). A Semantic Concordance.

Proceedings of the 3rd DARPA Workshop on Human Language Technology, 303-308.

Miller, GA., Beckwith, R., Fellbaum, C., Gross, D., Miller, K. (1990). WordNet: An on-

line lexical database. International Journal of Lexicography, 3, 235-244.

Ramshaw, L., Marcus, M. (1995). Text chunking using transformation-based learning.

Proceedings of the Third ACL Workshop on Very Large Corpora, 82-94.

Srinivas.,B. (1997). Performance evaluation of supertagging for partial parsing. In H.

Bunt and A. Nijholt, editors, Advances in probabilistic and Other Parsing Technologies,

Kluwer Academic Publishers, 203-220.

Stegman, R., Schulz, H., Hinrichs, E. (1998) Stylebook for the German Treebank in

VERBMOBIL. Manuskript Universität Tübingen.

Svartvik, J., Ed. (1990) The London Corpus of Spoken English: Description and

Research. Lund Studies in English 82. Lund University Press.

van Genabith, J., Way, A., Sadler, L. (1999) Semi-Automatic Generation of f-Structures

from Tree Banks. In M. Butt and T.H. King, editors, Proceedings of the LFG99

Conference, Manchester University, 19-21 July, CSLI Online, From Treebank Resources

to LFG f-Structures, Stanford, CA. http://www-csli.stanford.edu/publications/.

Voutilainen, A. (1995) A syntax-based part-of-speech analyzer Proceedings of the

European Chapter of the Association for Computational Linguistics, 157-164.

Xue, N., Chiou, F., Palmer, M. (2002). Building a large-scale annotated Chinese corpus.

Proceedings of the 19th. International Conference on Computational Linguistics

(COLING).

NANCY IDE